crossover from bose-einstein condensed molecules to...

13
Crossover from Bose-Einstein Condensed Molecules to Cooper Pairs by Using Feshbach Resonance Shu-Wei Chang Department of Electrical and Computer Engineering, University of Illinois at Urbana-Champaign, Urbana, IL 61801, USA (Dated: May 6, 2004) Abstract At low temperature, fermionic atoms can either condense into a Bose-Einstein condensate (BEC) by forming bosonic molecules or be loosely paired to form Cooper pairs described by Bardeen, Cooper, and Schrieffer’s (BCS) microscopic theory. Experimentalists have claimed the observation of the intermediate regime between BEC and BCS limits by tuning the scattering length with the aid of Feshbach resonance. In this report, we will discuss the idea of using Feshbach resonance to achieve the crossover. Some many-body formulation concerning this crossover will be briefly introduced, but not in detail. The experimental results will be our main concerns and addressed in this report. 1

Upload: others

Post on 27-Jan-2021

0 views

Category:

Documents


0 download

TRANSCRIPT

  • Crossover from Bose-Einstein Condensed Molecules to

    Cooper Pairs by Using Feshbach Resonance

    Shu-Wei Chang

    Department of Electrical and Computer Engineering,

    University of Illinois at Urbana-Champaign, Urbana, IL 61801, USA

    (Dated: May 6, 2004)

    Abstract

    At low temperature, fermionic atoms can either condense into a Bose-Einstein condensate (BEC)

    by forming bosonic molecules or be loosely paired to form Cooper pairs described by Bardeen,

    Cooper, and Schrieffer’s (BCS) microscopic theory. Experimentalists have claimed the observation

    of the intermediate regime between BEC and BCS limits by tuning the scattering length with the

    aid of Feshbach resonance. In this report, we will discuss the idea of using Feshbach resonance

    to achieve the crossover. Some many-body formulation concerning this crossover will be briefly

    introduced, but not in detail. The experimental results will be our main concerns and addressed

    in this report.

    1

  • I. INTRODUCTION

    After the successful demonstrations of Bose-Einstein condensation (BEC) of atomic

    gas,[1, 2], the cooling down of atomic Fermi gas has also been achieved recently.[3, 4] Unlike

    bosonic atoms, fermionic atoms cannot condense into a single state due to anti symmetriza-

    tion of the macroscopic wave function. However, two fermionic atoms can be converted into

    a bosonic molecule and condense into a BEC again.

    Whether two fermionic atoms can form a bosonic molecule is closely related to the two-

    body physics of the atoms. The sign of the s-wave scattering length between two identical

    fermionic atoms in different internal quantum states, called an open channel, determines

    whether the effective interaction in that channel between two fermions is repulsive or attrac-

    tive at low temperature.[5] The sign of the s-wave scattering length has a close connection

    with the near-by energy of the bound state in a closed channel of which threshold energy is

    higher than that of the open channel. For fermionic atoms with the same internal quantum

    states, the s-wave scattering cross section disappears due to anti-symmetrization of wave

    functions, and one has to consider scattering lengths due to, for example, p-wave scatter-

    ing. Experimentally, it will be more efficient to observe the change of the s-wave scattering

    length in an open channel characterized by the scattering of the atoms with different internal

    quantum numbers.[6, 7] However, if the physical environment favors the atoms of a single

    internal quantum state, issues on p-wave scattering such as the analogy of p-wave pairing

    in superconductivity then can be discussed.[8]

    For atoms, the s-wave scattering length can be controlled by mechanisms such as the

    interaction of magnetic moments with magnetic field[9] or Raman transition induced by

    lasers.[10, 11] In either case, the energy of the bound state in the closed channel is swept

    across the threshold energy of the open channel. It is possible to change the sign of the

    scattering length and thus the property of the effective interaction. For fermionic atoms,

    the controllability of the scattering length enables the formation of a correlated atom pair in

    one limit and a tightly-bound molecule in the other, as will be discussed in the next section.

    Experimentally, it is interesting to ask what happens to the pair of fermionic atoms at the

    crossover between the two limits. Theoretically, it can also provide an analogue test for one

    of the candidate theories of high temperature superconductivity, originated from Nozières

    and Schmitt-Rink by considering electron pairs resonant with the molecular state.[12]

    2

  • Eb

    E

    Eth

    Close Channel

    Open Channel

    Atomic Separation

    FIG. 1: Feshbach resonance of a quasi-bound state with the open channel. Eb is the resonant

    energy of the quasi-bound state. Eth is the threshold energy of the open channel. The blue part

    indicates the continuum of the open channel.

    II. FESHBACH RESONANCE

    Feshbach is first discussed in nuclear physics.[13] In this section, emphasis will be put on

    the Feshbach resonance using magnetic field. More details can be found in Ref. [5].

    For fermionic atoms, Feshbach resonance is the change of the scattering length in an open

    channel due to the coupling to a bound state in the closed channel. The state in the closed

    channel will be a true bound state if either there is no coupling to the open channel, or the

    energy of the state lies below the threshold energy of the open channel. On the other hand,

    if the state has an energy lying in the continuum of the open channel and exhibits a finite

    coupling to the open channel, it should be called a quasi-bound state since it is not a true

    bound state. A quasi-bound state exhibits some sort of energy resonance just as a bound

    state but is characterized by a finite decay width due to the coupling to the open channel.

    Fig. 1 shows the Feshbach resonance of a quasi-bound state coupling to the open channel.

    Write the general wave function |Ψ > as a superposition of the part in the open chan-

    nel |Ψ1 > and the part in the closed channel |Ψ2 >. The two parts are orthogonal to each

    other: < Ψ2|Ψ1 >= 0. The Hamiltonian describing both open and closed channels is de-

    noted as H. Also, write the projection operator to the open and closed channels as P1 and

    P2. We define various projected Hamiltonians as

    Hij = PiHPj, i, j = 1, 2 (1)

    3

  • After a close look, the effective Hamiltonian Heff describing the open channel is written as

    Heff |Ψ1 >=[

    H11 + H12(E − H22 + iδ)−1H21

    ]

    |Ψ1 >= E|Ψ1 > (2)

    where δ is a small number. We further write the projected Hamiltonian H11 as H11 = H0+U

    where H0 is the noninteracting part of the Hamiltonian H11. The interaction term Ueff in

    the effective Hamiltonian Heff is

    Ueff = U + H12(E − H22 + iδ)−1H21 (3)

    The effective s-wave scattering length is evaluated from equation (3) by Born’s approxima-

    tion for the zero momentum state |ϕ0 > of the open channel. The energy of the state |ϕ0 >

    is just the threshold energy Eth of the open channel. Denote a complete set of the eigenstates

    of the projected Hamiltonian H22 as {|φn >, n = 1, 2, ...} with eigenenergies {En}. Due to

    the fact that H12 = H†21 and the completeness relation 1 =

    n |φn >< φn| in the closed

    channel, the s-wave scattering length as of this open channel can be written as

    as =mr

    4π~2< ϕ0|Ueff |ϕ0 >

    =

    as,1︷ ︸︸ ︷mr

    4π~2< ϕ0|U |ϕ0 > +

    mr4π~2

    n

    | < φn|H21|ϕ0 > |2

    Eth − En + iδ

    =

    [

    as,1 +mr

    4π~2

    n,En 6=Eb

    | < φn|H21|ϕ0 > |2

    Eth − En + iδ

    ]

    ︸ ︷︷ ︸

    anr

    +mr

    4π~2| < φb|H21|ϕ0 > |

    2

    Eth − Eb + iδ(4)

    where as,1 denotes the s-wave scattering length in the open channel; |φb > is the (quasi-)

    bound state with energy Eb nearest to the threshold energy Eth; and ans is the nonresonant

    part of the s-wave scattering length. The contribution from the state |φb > is separated from

    the summation since it is the most significant. The denominator Eth −Eb in equation (4) is

    a function of the applied magnetic field B. Denote the internal quantum numbers labeling

    the two atoms in the open channel as α and β where α 6= β. Near a particular magnetic

    field B0, the denominator can be expanded as

    Eth − Eb ≈ (µb − µα − µβ)(B − B0) (5)

    where µb, µα, and µβ are the magnetic moments of the (quasi-) bound state and the states la-

    beled by internal quantum numbers α and β. If the nonresonant s-wave scattering length anr

    does not vary too much with the magnetic field, equation (4) can be written as

    4

  • B

    as

    anr

    B0

    FIG. 2: S-wave scattering length as a function of magnetic field.

    as = anr(1 +∆B

    B − B0 + iδB)

    ∆B =mr

    4π~2anr

    | < φb|H21|ϕ0 > |2

    µb − µα − µβ(6)

    where δB is an infinitesimal quantity. A typical s-wave scattering length as a function of

    the magnetic field is shown in Fig. 2 for anr > 0 and ∆ − B > 0. The scattering length

    diverges to plus and minus infinities as the magnetic field tends to B−0 and B+0 . Examples

    include the s-wave scattering length for |F = 9/2,mF = −9/2 > +|F = 9/2,mF = −7/2 >

    of 40K[6, 14, 15] and so on. Also, this scattering is an elastic one since the kinetic energy is

    not lost during the scattering.

    If the threshold energy Eth is slightly larger than Eb, the s-wave scattering length is

    positive. Although the effective interaction potential corresponding to this positive elastic

    scattering length is repulsive, two fermionic atoms can form a molecule when the magnetic

    field is swept downward across the magnetic field B0. In this case, it is possible to form

    BEC from these bosonic molecules. The energy difference between the two atoms and

    molecule is brought away by three-body processes instead of the s-wave two-body inelastic

    scattering which is believed to be absent.[8, 16–19] However, the molecules produced in this

    way are usually highly vibrationally excited. The deexcitation of these molecules will cause

    significant inelastic scattering, and the loss rate of these molecules is high.[6, 7, 20] On the

    other hand, for the negative scattering length (attractive interaction), atoms in the open

    channel cannot form a stable molecule described by a true bound state. They can interact

    via the quasi-bound state and form a correlated pair, or may be viewed as molecules with

    finite lifetime. This regime should be described by a generalized Bardeen, Cooper, and

    Schrieffer’s (BCS) theory instead of BEC. Typically, the atoms in this regime have lower

    loss rate than the molecules in the other side of the resonance. [6, 7, 20]

    5

  • III. ON THE MANY-BODY ASPECTS

    In atomic gases, the presence of quasi-bound state is easily modeled by adding a coupling

    term to a molecular state in the BCS Hamiltonian for fermionic atoms.[21–23]

    H =∑

    k

    ²k(a†k,↑ak,↑ + a

    †k,↓ak,↓) + Ubg

    k1+...+k4=0

    a†k1,↑a†k2,↓

    ak3,↓ak4,↑

    +g∑

    k,q

    (b†qaq/2+k,↑aq/2−k,↓ + bqa†

    q/2−k,↓a†

    q/2+k,↑) +∑

    q

    (ν + Eq)b†qbq (7)

    where a (b) and a† (b†) are the annihilation and creation operators of atoms (molecules); ²k

    and Eq are the the kinetic energies of atoms and molecules; ↑ and ↓ are pseudo spins indi-

    cating the two different internal quantum states; Ubg is the background attractive potential

    between atoms; µ is the bound-state energy relative to the threshold of the open channel

    due to the magnetic field; and g describes the exchange between atoms and molecules. This

    Hamiltonian describes the superfluidity of Fermi atoms in the BEC limit, BCS limit, and

    the intermediate regime. The change of the scattering length is not obvious in equation (7)

    but is implicitly included due to the coupling constant g. This model is a more rigorous

    picture than the two-body one in the previous section because the corresponding effective

    theory fails if the scattering length is large near the resonance.

    Instead of going through the mathematical details, we only mention a few results of key

    calculations. The detailed treatments of this Hamiltonian or its mutants for atomic Fermi

    gas can be found in literatures.[21–25]

    Fig. 3 shows the critical temperature, the chemical potential, and the fractions of atoms

    and molecules at the emergence of superfluidity as a function of the bound-state energy.[23]

    This result is obtained numerically by self-consistently solving the gap equation and the

    conservation of particle number.

    1 = (−Ubg +g2

    ν − 2µ)∑

    k

    tanh[β(k2/2m − µ)/2]

    2(k2/2m − µ)

    N = 2∑

    k

    1

    eβ(k2/2m−µ) + 1+ 2

    k

    1

    eβ(k2/4m+ν−µ) + 1−

    1

    β

    q

    ∂µln[1 − χ(q, iω)] (8)

    where µ is the chemical potential; N is the total particle number; and χ(q, iω) is a correction

    function describing the number fluctuations of atoms and molecules. In Fig. 3(a), the

    solid line shows the critical temperature calculated from equation (8). The dash line is

    a calculation by substituting equation (6) into Ubg and solving the BCS gap equation. The

    6

  • (a) (b) (c)

    FIG. 3: Physical quantities at the emergence of superfluidity as a function of the bound-state

    energy. (a) Critical temperature. Solid line is the result from equation (8). Dash-line is the result

    from the reduced BCS theory. (b) Chemical Potential. (c) Fractions of atoms and molecules. Solid

    line is the fraction of the fermions while the dot-dash line is the pair fraction. After Ref. [23]

    critical temperature calculated from simple BCS gap equation diverges at the resonance

    because the interaction energy Ubg calculated in this way diverges to minus infinity. The

    solid line matches well with the simple BCS theory when the bound-state energy is above

    the threshold. On the other hand, it approaches the critical temperature predicted by BEC

    theory at the other limit. As the bound-state energy is tuned downward from the BCS side

    toward the resonance, the critical temperature increases, which makes this model a candidate

    for high-temperature superconductivity. There is an optimized critical temperature close

    to the threshold energy. This critical temperature is around 0.26 TF , where TF is the

    Fermi temperature of the atoms without any molecules. Similarly in Fig. 3(b), the chemical

    potential is also a smooth connection between the two limits. In the BCS limit, the chemical

    potential approaches that of the Fermi gas at critical temperature. In the BEC limit, the

    chemical potential will tend to one half of the bound-state energy. This is an analogy to

    that the chemical potential of pure bosons approach their ground energy as BEC occurs.

    Fig. 3(c) shows the fractions of fermionic atoms and atom pairs. The fraction of fermions

    approach zero as the system goes deep into BEC limit while the pair fraction shows the

    opposite trend. The atom pairs resemble Cooper pairs above the resonance but are like

    molecules below the resonance.

    The calculation stated above is only related to the physical quantities when the super-

    fluidity phase occurs. The question on whether there is a transition between BEC and

    7

  • 0

    0.05

    0.1

    0.15

    0.2

    0.25

    0.3

    0.35

    0.4

    0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

    T/T

    F

    B-B0 (Gauss)

    BEC BCS

    TF=0.35 µK

    (a)

    0.1

    0.15

    0.2

    00.2

    0.40.6

    0

    0.03

    0.06

    T/TFB−B

    0 (Gauss)

    nm

    c/n

    m

    (b)

    FIG. 4: (a) The phase diagram of BCS-BEC transition. The upper dash line is the critical temper-

    ature calculated from simple BCS theory. The lower dash line is the BCS-BEC boundary. (b) The

    fraction of BEC condensate as a function of the magnetic field and temperature. After Ref. [26]

    BCS limits, i.e., a boundary between BEC and BCS regimes in the phase diagram shown in

    Fig. 4(a), is also interesting.[26] At zero temperature, write the pair wave function as the

    superposition of the open and closed-channel wave functions χm(x, x′) and χaa(x, x

    ′),

    Z(B)χm(x, x′)|closed > +

    1 − Z(B)χaa(x, x′)|open > (9)

    the question of BCS-BEC crossover becomes whether we can find a critical magnetic field

    Bc(T = 0) such that the parameter Z(B < Bc(T = 0)) = 1 and Z(B > Bc(T = 0)) = 0. At

    finite temperatures less than the critical temperature, the question becomes whether there is

    a critical magnetic field Bc(T ) below which molecules begin to condense into the molecular

    ground state, i.e., nm,BEC(B < Bc(T ), T )/nm(B < Bc(T ), T ) > 0. We first note that the

    critical temperature calculated in Ref. [23] has a different shape from that calculated in

    Ref. [26]. It may be a result of different approximations used in the calculations. Fig. 4(b)

    shows the BEC fraction as a fraction of the magnetic field and temperature. At a given

    temperature, there is a critical magnetic field Bc(T ) where the BEC condensate emerges.

    These critical magnetic fields Bc(T ) deviate from B0 predicted by simple two-body physics.

    A simple argument about these deviations is that molecules, though with a finite lifetime in

    the BCS regime, will emerge from atomic Fermi sea when the shifted ground state energy

    of molecules due to the coupling is roughly twice the chemical potential of the system.

    This early formation of the molecular condensate before the exact resonance makes the

    superfluidity at this crossover unusual.

    8

  • (a)

    3

    2

    1

    0

    N -

    2 [10

    -9]

    6004002000Time [ms]

    60

    40

    20

    0

    N -

    1 [10

    -6]

    a)

    b)

    (b)

    FIG. 5: (a) The scattering length of 40K. (b) The inelastic scattering rate of 6Li. The upper and

    lower figure are the inverse and inverse square of the population as a function of time, indicating

    the two-body and three-body processes. After Ref. [9, 14].

    IV. EXPERIMENTS ON BCS-BEC CROSSOVER

    Most of the experiments on BCS-BEC crossing are aimed at fermionic atoms 6Li and

    40K.[2, 4, 6, 7, 9, 14, 15, 17–19] For 6Li, two hyperfine states |1/2, 1/2 > and |1/2,−1/2 >

    are used while for 40K, two states |9/2,−7/2 > and |9/2,−9/2 > are used.

    Fig. 5(a) shows the elastic scattering length of 40K as a function of magnetic field.[14]

    The Feshbach resonance of 40K takes place at 224.18 G. The magnitude of the scattering

    length is obtained from the extraction of rethermalization time by heating the atomic gas

    when modulating the power of the optical trap.[9] The sign of the scattering length is

    obtained by measuring the mean-filed shift of the transition |9/2,−9/2 >−→ |9/2,−5/2 >

    .[15] The change of the scattering length is very significant as the magnetic field sweeps

    across the resonance frequency, indicating the ability to control the effective interaction

    between 40K atoms. Fig. 5(b) shows the inelastic loss rate of 6Li at 680 G.[17] There are two

    Feshbach resonance peaks for 6Li–550 G and 822 G. The resonance at 822 G has a very broad

    lineshape, and thus the inelastic scattering at 680 G is believed to be greatly influenced by

    this resonance. The upper plot of Fig. 5(b) shows the inverse of the number of particles

    remaining in the trap as a function of time. The nearly linear behavior seems to imply

    that the two-body inelastic process dominates in the time scale observed. The lower plot

    shows the inverse square of the population as a function of time. The three-body process will

    dominate over the two-body process only when the population is high. The small linear part

    at time close to zero thus indicates the three-body process. However, the inelastic two-body

    9

  • 201 202 2030

    20

    40

    60

    80

    100re

    main

    ing

    ato

    ms

    (%)

    B (gauss)

    (a)

    Ato

    mic

    Fra

    ctio

    n

    Magnetic Field [G]

    1.0

    0.5

    0.0

    900850800750

    (b)

    FIG. 6: (a) The fraction of 40K atoms as a function of the magnetic field. (b) The fraction of 6Li

    atoms as a function of the magnetic field. Cross: initial ramp-down speed=30 G/µs; circle: linear

    ramp-down at 100 G/ms; triangle: linear ramp-down at12.5 G/ms. After Ref. [7, 19].

    loss is believed to be suppressed. In Ref. [17], the author thus attributed the observed loss

    rate to the three-body process simultaneously accompanying by the temperature change.

    Fig. 6(a) and (b) show the fractions of fermionic atoms 40K at 224.18 G[19] and 6Li at

    822 G[7] as a function of the magnetic field. In either case, the measurement begins from

    the BEC side where bound molecules exist. The magnetic field is first swept up to a given

    value. If the given magnetic field exceeds the resonant magnetic field, the molecules will

    dissociate. During the sweeping which takes about more than ten milliseconds, the molecular

    or atomic gas will expand so that the density of the gas is diluted. This action is performed

    to avoid the unwanted interference from many-body effects and the reformation of molecules

    at successive stage.[6, 7] After the expansion, the magnetic field is fast ramped down to zero.

    The fraction of the atoms is then obtained at zero field by optical image method which only

    detects the atoms because the transition of molecules does not match the frequency of the

    incident light.[6, 14, 27] From Fig. 6(a) and (b), the fraction of atoms greatly increase as the

    given magnetic field goes above the resonance, indicating the absence of molecules. Also, in

    Fig. 6(b), the fraction of atoms corresponding to different speeds of ramping the magnetic

    field down to zero are shown. This ramp-down speed has to be carefully chosen because the

    resonant field B0 obtained in this way will shift due to the ramp-down speed.[7] However,

    the ramp-down speeds used in the experiments do not significantly change the resonant field.

    Fig. 7(a) and (b) show the fraction of the condensate of pair atoms (This term should

    10

  • -0.5 0 0.5 1.00

    0.05

    0.10

    0.15

    -9 -6 -3 0 3 16 18198

    200

    202

    204

    B(g

    auss

    )

    t (ms)N

    0/N

    ∆B (gauss)

    (a)

    Co

    nd

    en

    sa

    te F

    ractio

    n

    Magnetic Field [G]

    0.8

    0.4

    0.0

    1000800600

    (b)

    FIG. 7: (a) The decay of 40K2 condensate fraction. Circle: 2 ms hold time; triangle: 30 ms hold

    time (b) The decay of 6Li2 condensate fraction. Cross: 2 ms hold time; square: 100 ms hold time;

    circle: 10 s hold time. After Ref. [6, 7].

    stand for BEC molecules in the BCS side although its real meaning is a bit ambiguous in

    these papers.) as a function of magnetic field for 40K[6] and 6Li.[7] Unlike Fig. 6(a) and (b),

    the initial magnetic field is set at the BCS side and than brought slowly to a given magnetic

    field to avoid the enhanced loss out of the optical trap due to rapid sweeping. No expansion

    of atomic gas takes place at this stage. Furthermore, sometimes it is even necessary to in

    crease optical power to enhance the confinement of the atomic gas.[7] The atomic gas is

    held there for a period of time which is variable in the experiments. After the hold period,

    the trap is turned off to allow the gas to expand. Simultaneously[6] or immediately after

    the expansion,[7] the magnetic field is soon turned off. Unlike the experiments mentioned

    previously, the atom in the gas will be adiabatically turned into molecules due to high

    density. The fraction of molecular BEC condensate, which is assumed to be invariant before

    and after the turn-off by some subtle argument, is then measured by mapping out the zero

    momentum component[7] or Thomas-Fermi profile of the molecules.[6]

    In Fig 7 (a) and (b), the fraction of the condensate decays as the hold time increases. As

    mentioned in section II, the formation of molecules by sweeping the magnetic field across the

    resonance creates vibrationlly-excited molecules. The deexcitation of these molecules will

    cause the inelastic scattering and quench the fraction od the BEC condensate. The longer

    the hold time is, the more thorough the quenching is. It accounts why significant decay is

    observed at longer hold time. Close to the resonance and in the BCS side, the condensate

    decays slowly while it decays faster farther away from the resonance in the BEC side because

    the molecules are more tightly-bound. The extra bound energy has to be converted into

    11

  • 0

    0.05

    0.10

    0.15

    0.20

    -0.4-0.2

    00.2

    0.40.6

    0.8

    -0.05

    0

    0.05

    0.10

    0.15

    N /N0

    T/TF

    ∆B (gauss)

    (a)

    0

    0.2

    0.4

    0.6

    Condensate

    Fra

    ction a

    t 820 G

    1000900800700Magnetic Field [G]

    0.5

    0.3

    0.1

    0.7

    0.8

    0.2

    0.05

    T/TF

    (b)

    FIG. 8: The fractions of (a) 40K2 BEC condensate and (b)6Li2 BEC condensate as a function of

    the magnetic field and temperature. After Ref. [6, 7].

    kinetic energy and reduces the condensate fraction.[28]

    The argument that the condensate fraction measured in this way is just the condensate

    fraction before the turn-off of the magnetic field is subtle. This argument will lead to the

    conjecture that the BEC is present in the BCS side.[26] The argument is based on two

    assumptions. The first one is that no BEC condensate is created when the magnetic field

    drops to zero.[14] The second one is that slightly-correlated atoms tend to randomly pick up

    a neighboring atom to form a molecule. Molecules formed in this way have finite momenta

    and do not belong to the condensate. Therefore, the original condensate will be directly

    projected to the condensate at zero magnetic field.[7] The condensate fraction as a function

    of magnetic field and temperature for a given number of atoms is also measured.[6, 7] Fig. 8

    (a) and (b) show the corresponding three-dimensional and contour plots for 40K2 and6Li2

    molecules. Fig. 8 (a) can be directly compared with Fig. 4 (b) since this figure is an attempt

    to fit the experimental data.

    V. CONCLUSION

    We have briefly surveyed the recent experiments on BEC-BCS crossover. Instead of going

    into complicated calculations of many-body dynamics, we only look into those related to

    thermal equilibrium. The interesting point of the cold fermionic atoms is the capability of

    tuning the scattering length by magnetic field. BEC and BCS types of superfluidity can

    thus be achieved in the same system by tuning this controllable parameter. Further, an

    interesting regime of superfluidity between BEC and BCS has been opened. More future

    work has to be devoted to explore the properties of the superfluidity phase in the regime.

    12

  • [1] M. H. Anderson et al., Science 269, 198 (1995).

    [2] K. B. Davis et al., Phys. Rev. Lett. 75, 3969 (1995).

    [3] F. A. van Abeelen, B. J. Verhaar, and A. J. Moerdijk, Phys. Rev. A 55, 4377 (1997).

    [4] B. DeMarco and D. S. Jin, Science 285, 1703 (1999).

    [5] C. Pethick and H. Smith, Bose-Einstein Condensation in Dilute Gases (Cambridge University

    Press, Cambridge, UK, 2002), 1st ed.

    [6] C. Regal, M. Greiner, and D. Jin, Phys. Rev. Lett. 92, 040403 (2004).

    [7] M. W. Zwierlein et al., Phys. Rev. Lett. 92, 120403 (2004).

    [8] J. L. Bohn, Phys. Rev. A 61, 053409 (1999).

    [9] T. Loftus et al., Phys. Rev. Lett. 88, 173201 (2002).

    [10] F. K. Fatemi, K. M. Jones, and P. D. Lett, Phys. Rev. Lett. 85, 4462 (2000).

    [11] S. Kokkelmans, H. M. J. Vissers, and B. J. Verhaar, Phys. Rev. Lett. 63, 031601(R) (2001).

    [12] P. Nozières and S. Schmitt-Rink, J. Low Temp. Phys. 59, 195 (1985).

    [13] H. Feshbach, Ann. Phys. 19, 287 (1962).

    [14] C. A. Regal, C. Ticknor, J. L. Bohn, and D. S. Jin, Nature (London) 424, 47 (2003).

    [15] C. A. Regal and D. S. Jin, Phys. Rev. Lett. 90, 230404 (2003).

    [16] V. A. Yurovsky and A. Ben-Reuven, Phys. Rev. A 60, R765 (1999).

    [17] K. Dieckmann et al., Phys. Rev. Lett. 89, 203201 (2002).

    [18] C. A. Regal, C. Ticknor, J. L. Bohn, and D. S. Jin, Phys. rev. Lett. 90, 053201 (2003).

    [19] C. A. Regal, M. Greiner, and D. S. Jin, Phys. Rev. Lett. 92, 083201 (2004).

    [20] K. E. Strecker, G. B. Partridge, and R. G. Hulet, Phys. Rev. Lett. 91, 080406 (2003).

    [21] M. Holland et al., Phys. Rev. Lett. 87, 120406 (2001).

    [22] Y. Ohashi and A. Griffin, Phys. Rev. Lett. 89, 130402 (2002).

    [23] J. N. Milstein, S. Kokkelmans, and M. J. Holland, Phys. Rev. A 66, 043604 (2002).

    [24] M. Mackie, K.-A. Suominen, and J. Javanainen, Phys. Rev. Lett. 89, 180403 (2002).

    [25] P. Naidon, Grancoise, and Masnou-Seeuws, Phys. Rev. A 68, 033612 (2003).

    [26] G. M. Falco and H. T. C. Stoof, Phys. Rev. Lett. 92, 130401 (2004).

    [27] M. W. Zwierlein et al., Phys. Rev. Lett. 91, 250401 (2003).

    [28] J. Cubizolles et al., Phys. Rev. Lett. 91, 240401 (2003).

    13